More stories

  • in

    Nitrogen addition decreases methane uptake caused by methanotroph and methanogen imbalances in a Moso bamboo forest

    1.
    Ni, X. & Groffman, P. M. Declines in methane uptake in forest soils. Proc. Natl. Acad. Sci. USA 115, 8587–8590 (2018).
    CAS  PubMed  Article  Google Scholar 
    2.
    IPCC. Climate change 2013: the physical science basis Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge University Press, 2013).

    3.
    Kirschke, S. et al. Three decades of global methane sources and sinks. Nat. Geosci. 6, 813–823 (2013).
    ADS  CAS  Article  Google Scholar 

    4.
    Turner, A. J., Frankenberg, C. & Kort, E. A. Interpreting contemporary trends in atmospheric methane. Proc. Natl. Acad. Sci. USA 116, 2805–2813 (2019).
    CAS  PubMed  Article  Google Scholar 

    5.
    Tate, K. R. Soil methane oxidation and land-use change–from process to mitigation. Soil Biol. Biochem. 80, 260–272 (2015).
    CAS  Article  Google Scholar 

    6.
    Thauer, R. K., Anne-Kristin, K., Henning, S., Wolfgang, B. & Reiner, H. Methanogenic archaea: ecologically relevant differences in energy conservation. Nat. Rev. Microbiol. 6, 579–591 (2008).
    CAS  PubMed  Article  Google Scholar 

    7.
    Banger, K., Tian, H. & Lu, C. Do nitrogen fertilizers stimulate or inhibit methane emissions from rice fields?. Glob. Change Biol. 18, 3259–3267 (2012).
    ADS  Article  Google Scholar 

    8.
    Murase, J. & Kimura, M. Methane production and its fate in paddy fields. IV. Sources of microorganisms and substrates responsible for anaerobic CH4 oxidation in subsoil. Soil Sci. Plant Nutr. 40, 57–61 (1994).
    CAS  Article  Google Scholar 

    9.
    Zhang, M., Huang, J., Sun, S., Rehman, M. & He, S. Depth-specific distribution and significance of nitrite-dependent anaerobic methane oxidation process in tidal flow constructed wetlands used for treating river water. Sci. Total Environ. 716, 107354 (2020).
    Google Scholar 

    10.
    Yu, X. et al. Sonneratia apetala introduction alters methane cycling microbial communities and increases methane emissions in mangrove ecosystems. Soil Biol. Biochem. 144, 107775 (2020).
    CAS  Article  Google Scholar 

    11.
    Hanson, R. S. & Hanson, T. E. Methanotrophic bacteria. Microbiol. Rev. 60, 439–471 (1996).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Knief, C. Diversity and habitat preferences of cultivated and uncultivated aerobic methanotrophic bacteria evaluated based on pmoA as molecular marker. Front. Microbiol. 6, 1346 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    13.
    Dunfield, P., Knowles, R., Dumont, R. & Moore, T. R. Methane production and consumption in temperate and subarctic peat soils: response to temperature and pH. Soil Biol. Biochem. 25, 321–326 (1993).
    CAS  Article  Google Scholar 

    14.
    Mer, J. L. & Roger, P. Production, oxidation, emission and consumption of methane by soils: a review. Eur. J. Soil Biol. 37, 25–50 (2001).
    Article  Google Scholar 

    15.
    Aronson, E. L., Dubinsky, E. A. & Helliker, B. R. Effects of nitrogen addition on soil microbial diversity and methane cycling capacity depend on drainage conditions in a pine forest soil. Soil Biol. Biochem. 62, 119–128 (2013).
    CAS  Article  Google Scholar 

    16.
    Bodelier, P. L. E. & Laanbroek, H. J. Nitrogen as a regulatory factor of methane oxidation in soils and sediments. FEMS Microbiol. Ecol. 47, 265–277 (2004).
    CAS  PubMed  Article  Google Scholar 

    17.
    Galloway, J. N. et al. Transformation of the nitrogen cycle: recent trends, questions, and potential solutions. Science 320, 889–892 (2008).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    18.
    Liu, L. & Greaver, T. L. A review of nitrogen enrichment effects on three biogenic GHGs: the CO2 sink may be largely offset by stimulated N2O and CH4 emission. Ecol. Lett. 12, 1103–1117 (2009).
    CAS  PubMed  Article  Google Scholar 

    19.
    Fowler, D., Coyle, M., Skiba, U., Sutton, M. A. & Voss, M. The global nitrogen cycle in the twenty-first century. Philos. Trans. R. Soc. B. 368, 20130164 (2013).
    Article  CAS  Google Scholar 

    20.
    Reay, D. S., Dentener, F., Smith, P., Grace, J. & Feely, R. A. Global nitrogen deposition and carbon sinks. Nat. Geosci. 1, 430–437 (2008).
    ADS  CAS  Article  Google Scholar 

    21.
    Ackerman, D., Millet, D. B. & Chen, X. Global estimates of inorganic nitrogen deposition across four decades. Glob. Biogeochem. Cycles 33, 100–107 (2019).
    ADS  CAS  Article  Google Scholar 

    22.
    Liu, X. et al. Enhanced nitrogen deposition over China. Nature 494, 459–462 (2013).
    ADS  CAS  PubMed  Article  Google Scholar 

    23.
    Li, Q. et al. Nitrogen depositions increase soil respiration and decrease temperature sensitivity in a Moso bamboo forest. Agric. For. Meteorol. 268, 48–54 (2019).
    ADS  Article  Google Scholar 

    24.
    Steudler, P. A., Bowden, R. D., Melillo, J. M. & Aber, J. D. Influence of nitrogen fertilization on methane uptake in temperate forest soils. Nature 341, 314–316 (1989).
    ADS  Article  Google Scholar 

    25.
    Hütsch, B. W., Webster, C. P. & Powlson, D. S. Methane oxidation in soil as affected by land use, soil pH and N fertilization. Soil Biol. Biochem. 26, 1613–1622 (1994).
    Article  Google Scholar 

    26.
    Bodelier, P. L. E., Roslev, P., Henckel, T. & Frenzel, P. Stimulation by ammonium-based fertilizers of methane oxidation in soil around rice roots. Nature 403, 421–424 (2000).
    ADS  CAS  PubMed  Article  Google Scholar 

    27.
    Kruger, M. & Frenzel, P. Effects of N-fertilisation on CH4 oxidation and production, and consequences for CH4 emissions from microcosms and rice fields. Glob. Change Biol. 9, 773–784 (2003).
    ADS  Article  Google Scholar 

    28.
    Delgado, J. A. & Mosier, A. R. Mitigation alternatives to decrease nitrous oxides emissions and urea-nitrogen loss and their effect on methane flux. J. Environ. Qual. 25, 1105–1111 (1996).
    CAS  Article  Google Scholar 

    29.
    Shang, Q. et al. Net annual global warming potential and greenhouse gas intensity in Chinese double rice-cropping systems: a 3-year field measurement in long-term fertilizer experiments. Glob. Change Biol. 17, 2196–2210 (2011).
    ADS  Article  Google Scholar 

    30.
    Cai, Z. et al. Methane and nitrous oxide emissions from rice paddy fields as affected by nitrogen fertilizers and water management. Plant Soil 196, 7–14 (1997).
    CAS  Article  Google Scholar 

    31.
    Malghani, S., Reim, A., Fischer, J. V., Conrad, R. & Trumbore, S. E. Soil methanotroph abundance and community composition are not influenced by substrate availability in laboratory incubations. Soil Biol. Biochem. 101, 184–194 (2016).
    CAS  Article  Google Scholar 

    32.
    Schnyder, E., Bodelier, P. L. E., Hartmann, M., Henneberger, R. & Niklaus, P. A. Positive diversity-functioning relationships in model communities of methanotrophic bacteria. Ecology 99, 714–723 (2018).
    PubMed  Article  Google Scholar 

    33.
    Wang, C., Liu, D. & Bai, E. Decreasing soil microbial diversity is associated with decreasing microbial biomass under nitrogen addition. Soil Biol. Biochem. 120, 126–133 (2018).
    CAS  Article  Google Scholar 

    34.
    Shrestha, M., Shrestha, P. M., Frenzel, P. & Conrad, R. Effect of nitrogen fertilization on methane oxidation, abundance, community structure, and gene expression of methanotrophs in the rice rhizosphere. ISME J. 4, 1545–1556 (2010).
    CAS  PubMed  Article  Google Scholar 

    35.
    Liu, H. et al. Responses of soil methanogens, methanotrophs, and methane fluxes to land-use conversion and fertilization in a hilly red soil region of southern China. Environ. Sci. Pollut. Res. 24, 8731–8743 (2017).
    CAS  Article  Google Scholar 

    36.
    Bao, Q., Ding, L. J., Huang, Y. & Xiao, K. Effect of rice straw and/or nitrogen fertiliser inputs on methanogenic archaeal and denitrifying communities in a typical rice paddy soil. Earth Environ. Sci. Trans. R. Soc. Edinb. 109, 375–386 (2019).
    CAS  Google Scholar 

    37.
    Ho, A. et al. The more, the merrier: heterotroph richness stimulates methanotrophic activity. ISME J. 8, 1945–1948 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    38.
    Dan, H. et al. The response of methanotrophs to additions of either ammonium, nitrate or urea in alpine swamp meadow soil as revealed by stable isotope probing. FEMS Microbiol. Ecol. 7, fiz077 (2019).
    Google Scholar 

    39.
    Zhang, D., Mo, L., Chen, X., Zhang, L. & Xu, X. Effect of nitrogen addition on methanotrophs in temperate forest soil. Acta Ecol. Sin. 37, 8254–8263 (2017).
    Google Scholar 

    40.
    Mohanty, S. R., Bodelier, P. L. E., Floris, V. & Conrad, R. Differential effects of nitrogenous fertilizers on methane-consuming microbes in rice field and forest soils. Appl. Environ. Microbiol. 72, 1346–1354 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    41.
    Hu, A. & Lu, Y. The differential effects of ammonium and nitrate on methanotrophs in rice field soil. Soil Biol. Biochem. 85, 31–38 (2015).
    CAS  Article  Google Scholar 

    42.
    Shrestha, P. M. et al. Linking activity, composition and seasonal dynamics of atmospheric methane oxidizers in a meadow soil. ISME J. 6, 1115–1126 (2012).
    CAS  PubMed  Article  Google Scholar 

    43.
    Jang, I., Lee, S., Zoh, K. D. & Kang, H. Methane concentrations and methanotrophic community structure influence the response of soil methane oxidation to nitrogen content in a temperate forest. Soil Biol. Biochem. 43, 620–627 (2011).
    CAS  Article  Google Scholar 

    44.
    Song, X., Chen, X., Zhou, G., Jiang, H. & Peng, C. Observed high and persistent carbon uptake by Moso bamboo forests and its response to environmental drivers. Agric. For. Meteorol. 247, 467–475 (2017).
    ADS  Article  Google Scholar 

    45.
    Song, X. et al. Carbon sequestration by Chinese bamboo forests, and their ecological benefits: assessment of potential, problems, and future challenges. Environ. Rev. 19, 418–428 (2011).
    CAS  Article  Google Scholar 

    46.
    Jia, Y. et al. Spatial and decadal variations in inorganic nitrogen wet deposition in China induced by human activity. Sci. Rep. 4, 3763 (2014).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    47.
    Song, X., Zhou, G., Gu, H. & Qi, L. Management practices amplify the effects of N deposition on leaf litter decomposition of the Moso bamboo forest. Plant Soil 395, 391–400 (2015).
    CAS  Article  Google Scholar 

    48.
    Mo, J., Fang, Y., Xu, G., Li, D. & Xue, J. The short-term responses of soil CO2 emission and CH4 uptake to simulated N deposition in nursery and forests of Dinghushan in subtropical China. Acta Ecol. Sin. 25, 682–690 (2005).
    CAS  Google Scholar 

    49.
    Zhang, W. et al. Methane uptake responses to nitrogen deposition in three tropical forests in southern China. J. Geophys. Res. 113, D11116 (2008).
    ADS  Article  CAS  Google Scholar 

    50.
    Song, X. et al. Nitrogen addition increased CO2 uptake more than non-CO2 greenhouse gases emissions in a Moso bamboo forest. Sci. Adv. 6, eaaw5790 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    51.
    Knief, C., Lipski, A. & Dunfield, P. F. Diversity and activity of methanotrophic bacteria in different upland soils. Appl. Environ. Microbiol. 69, 6703–6714 (2003).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    52.
    Wang, M., Xu, X., Wang, W., Wang, G. & Su, C. Effects of slag and biochar amendments on methanogenic community structures in paddy fields. Acta Ecol. Sin. 38, 2816–2818 (2018).
    Article  Google Scholar 

    53.
    Zeikus, J. G. Biology of methanogenic bacteria. Bacteriol. Rev. 41, 514–541 (1977).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    54.
    Täumer, J. et al. Divergent drivers of the microbial methane sink in temperate forest and grassland soils. Glob. Change Biol. 27, 929–940 (2021).

    55.
    Pratscher, J., Vollmers, J., Wiegand, S., Dumont, M. G. & Kaster, A. K. Unravelling the identity, metabolic potential and global biogeography of the atmospheric methane-oxidizing upland soil cluster α. Environ. Microbiol. 20(3), 1016–1029 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    56.
    Knief, C. Diversity and habitat preferences of cultivated and uncultivated aerobic methanotrophic bacteria evaluated based on pmoA as molecular marker. Front. Microbiol. 6, 487 (2015).
    Article  Google Scholar 

    57.
    Deng, Y. et al. Upland soil cluster gamma dominates methanotrophic communities in upland grassland soils. Sci. Total Environ. 670, 826–836 (2019).
    ADS  CAS  PubMed  Article  Google Scholar 

    58.
    Henckel, T., Friedrich, M. & Conrad, R. Molecular analyses of the methane-oxidizing microbial community in rice field soil by targeting the genes of the 16S rRNA, particulate methane monooxygenase, and methanol dehydrogenase. Appl. Environ. Microbiol. 65, 1980–1990 (1999).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    59.
    Lieberman, R. L. & Rosenzweig, A. C. Biological methane oxidation: regulation, biochemistry, and active site structure of particulate methane monooxygenase. Crit. Rev. Biochem. Mol. Biol. 39, 147–164 (2004).
    CAS  PubMed  Article  Google Scholar 

    60.
    Freitag, T. E. & Prosser, J. I. Correlation of methane production and functional gene transcriptional activity in a peat soil. Appl. Environ. Microbiol. 75, 6679–6687 (2009).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    61.
    Thauer, R. K. Biochemistry of methanogenesis: a tribute to Marjory Stephenson: 1998 Marjory Stephenson prize lecture. Microbiology 144, 2377–2406 (1998).
    CAS  PubMed  Article  Google Scholar 

    62.
    Schnell, S. & King, G. M. Mechanistic analysis of ammonium inhibition of atmospheric methane consumption in forest soils. Appl. Environ. Microbiol. 60, 3514–3521 (1994).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    63.
    Chao, A. Nonparametric estimation of the number of classes in a population. Scand. J. Stat. 11, 265–270 (1984).
    MathSciNet  Google Scholar 

    64.
    Shannon, C. E. A. mathematical theory of communication. Bell Syst. Tech. J. 27, 379–423 (1948).
    MathSciNet  MATH  Article  Google Scholar 

    65.
    Li, Q. et al. Biochar amendment decreases soil microbial biomass and increases bacterial diversity in Moso bamboo (Phyllostachys edulis) plantations under simulated nitrogen deposition. Environ. Res. Lett. 13, 044029 (2018).
    ADS  Article  CAS  Google Scholar 

    66.
    Li, Q., Song, X., Gu, H. & Gao, F. Nitrogen deposition and management practices increase soil microbial biomass carbon but decrease diversity in Moso bamboo plantations. Sci. Rep. 6, 28235 (2016).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    67.
    Frey, S. D., Knorr, M., Parrent, J. L. & Simpson, R. T. Chronic nitrogen enrichment affects the structure and function of the soil microbial community in temperate hardwood and pine forests. For. Ecol. Manag. 196, 159–171 (2004).
    Article  Google Scholar 

    68.
    Lin, Y. et al. Long-term application of lime or pig manure rather than plant residues suppressed diazotroph abundance and diversity and altered community structure in an acidic ultisol. Soil Biol. Biochem. 123, 218–228 (2018).
    CAS  Article  Google Scholar 

    69.
    Rousk, J. et al. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 4, 1340–1351 (2010).
    PubMed  Article  Google Scholar 

    70.
    Zhou, X., Guo, Z., Chen, C. & Jia, Z. Soil microbial community structure and diversity are largely influenced by soil pH and nutrient quality in 78-year-old tree plantations. Biogeosciences 14, 2101–2111 (2017).
    ADS  CAS  Article  Google Scholar 

    71.
    Nicol, G. W., Leininger, S., Schleper, C. & Prosser, J. I. The influence of soil pH on the diversity, abundance and transcriptional activity of ammonia oxidizing archaea and bacteria. Environ. Microbiol. 10, 2966–2978 (2008).
    CAS  PubMed  Article  Google Scholar 

    72.
    Vitousek, P. M. et al. Technical report: human alteration of the global nitrogen cycle: sources and consequences. Ecol. Appl. 7, 737 (1997).
    Google Scholar 

    73.
    Treseder, K. K. Nitrogen additions and microbial biomass: a meta-analysis of ecosystem studies. Ecol. Lett. 11, 1111–1120 (2008).
    PubMed  Article  Google Scholar 

    74.
    Serna-Chavez, H. M. & Bodegom, P. M. V. Global drivers and patterns of microbial abundance in soil. Glob. Ecol. Biogeogr. 22, 1162–1172 (2013).
    Article  Google Scholar 

    75.
    Rosso, L., Lobry, J. R., Bajard, S. & Flandrois, J. P. Convenient model to describe the combined effects of temperature and pH on microbial growth. Appl. Environ. Microbiol. 61, 610–616 (1995).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    76.
    Högberg, M. N., Högberg, P. & Myrold, D. D. Is microbial community composition in boreal forest soils determined by pH, C-to-N ratio, the trees, or all three?. Oecologia 150, 590–601 (2007).
    ADS  PubMed  Article  Google Scholar 

    77.
    Sterner, R. W. & Elser, J. J. Ecological Stoichiometry: The Biology of Elements from Molecules to the Biosphere (Princeton University Press, Princeton, 2002).
    Google Scholar 

    78.
    Kolb, S. The quest for atmospheric methane oxidizers in forest soils. Environ. Microbiol. Rep. 1, 336–346 (2009).
    CAS  PubMed  Article  Google Scholar 

    79.
    Topp, E. & Pettey, E. Soils as sources and sinks for atmospheric methane. Can. J. Soil Sci. 77, 167–177 (1997).
    CAS  Article  Google Scholar 

    80.
    Bender, M. & Conrad, R. Effect of CH4 concentrations and soil conditions on the induction of CH4 oxidation activity. Soil Biol. Biochem. 27, 1517–1527 (1995).
    CAS  Article  Google Scholar 

    81.
    Kolb, S., Knief, C., Dunfield, P. F. & Conrad, R. Abundance and activity of uncultured methanotrophic bacteria involved in the consumption of atmospheric methane in two forest soils. Environ. Microbiol. 7(8), 1150–1161 (2005).
    CAS  PubMed  Article  Google Scholar 

    82.
    Degelmann, D. M., Borken, W., Drake, H. L. & Kolb, S. Different atmospheric methane-oxidizing communities in European Beech and Norway Spruce Soils. Appl. Environ. Microbiol. 76(10), 3228–3235 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    83.
    Li, S., Yu, Y. & He, S. Summary of research on dissolved organic carbon (DOC). Soil Environ. Sci. 11, 422–429 (2002).
    Google Scholar 

    84.
    Zhang, R. et al. Nitrogen deposition enhances photosynthesis in Moso bamboo but increases susceptibility to other stress factors. Front. Plant Sci. 8, 1975 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    85.
    Wan, X. et al. Soil C:N ratio is the major determinant of soil microbial community structure in subtropical coniferous and broadleaf forest plantations. Plant Soil 387, 103–116 (2015).
    CAS  Article  Google Scholar 

    86.
    Demoling, F., Figueroa, D. & Bååth, E. Comparison of factors limiting bacterial growth in different soils. Soil Biol. Biochem. 39, 485–2495 (2007).
    Article  CAS  Google Scholar 

    87.
    Aronson, E. L. & Helliker, B. R. Methane flux in non-wetland soils in response to nitrogen addition: a meta-analysis. Ecology 91, 3242–3251 (2010).
    CAS  Article  Google Scholar 

    88.
    Cheng, S. et al. The primary factors controlling methane uptake from forest soils and their responses to increased atmospheric nitrogen deposition: a review. Acta Ecol. Sin. 32, 4914–4923 (2012).
    ADS  CAS  Article  Google Scholar 

    89.
    Fierer, N. et al. Comparative metagenomic, phylogenetic and physiological analyses of soil microbial communities across nitrogen gradients. ISME J. 6, 1007–1017 (2012).
    CAS  PubMed  Article  Google Scholar 

    90.
    Ramirez, K. S., Craine, J. M. & Fierer, N. Consistent effects of nitrogen amendments on soil microbial communities and processes across biomes. Glob. Change Biol. 18, 1918–1927 (2012).
    ADS  Article  Google Scholar 

    91.
    Song, X., Li, Q. & Gu, H. Effect of nitrogen deposition and management practices on fine root decomposition in Moso bamboo plantations. Plant Soil 410, 207–215 (2017).
    CAS  Article  Google Scholar 

    92.
    Vance, E. D., Brookes, P. C. & Jenkinson, D. S. An extraction method for measuring soil microbial biomass C. Soil Biol. Biochem. 19, 703–707 (1987).
    CAS  Article  Google Scholar 

    93.
    Li, Y. et al. Biochar reduces soil heterotrophic respiration in a subtropical plantation through increasing soil organic carbon recalcitrancy and decreasing carbon-degrading microbial activity. Soil Biol. Biochem. 122, 173–185 (2018).
    CAS  Article  Google Scholar 

    94.
    Lu, R. Methods for Soil Agro-chemistry Analysis (China Agricultural Science and Technology Press, Beijing, 2000).
    Google Scholar 

    95.
    Bourne, D. G., Mcdonald, I. R. & Murrell, J. C. Comparison of pmoA PCR primer sets as tools for investigating methanotroph diversity in three Danish soils. Appl. Environ. Microbiol. 67, 3802 (2001).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    96.
    Angel, R., Claus, P. & Conrad, R. Methanogenic archaea are globally ubiquitous in aerated soils and become active under wet anoxic conditions. ISME J. 6, 847–862 (2011).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    97.
    Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7, 335–336 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    98.
    Wang, Q. et al. Ecological patterns of nifH genes in four terrestrial climatic zones explored with targeted metagenomics using FrameBot, a new informatics tool. mBio 4, e00592-e613 (2013).
    PubMed  PubMed Central  Google Scholar 

    99.
    Kou, Y. et al. Scale-dependent key drivers controlling methane oxidation potential in Chinese grassland soils. Soil Biol. Biochem. 111, 104–114 (2017).
    CAS  Article  Google Scholar 

    100.
    Kou, Y. et al. Climate and soil parameters are more important than denitrifier abundances in controlling potential denitrification rates in Chinese grassland soils. Sci. Total Environ. 669, 62–69 (2019).
    ADS  CAS  PubMed  Article  Google Scholar 

    101.
    Wei, H. et al. Contrasting soil bacterial community, diversity, and function in two forests in China. Front. Microbiol. 9, 1693 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    102.
    Liu, W. et al. Critical transition of soil bacterial diversity and composition triggered by nitrogen enrichment. Ecology 101, e03053 (2020).
    PubMed  Google Scholar 

    103.
    Tang, X., Liu, S., Zhou, G., Zhang, D. & Zhou, C. Soil-atmospheric exchange of CO2, CH4, and N2O in three subtropical forest ecosystems in southern China. Glob. Change Biol. 12, 546–560 (2006).
    ADS  Article  Google Scholar  More

  • in

    The flight of the hornbill: drift and diffusion in arboreal avian movement

    A mathematical model to simulate movement
    For ‘attracting features’, such as nesting or roosting sites, we employ potential terms that are logarithmic in distance. Logarithmic potentials have been employed in diffusion models7 such as those involving long-range interactions8. The forces due to these are inversely proportional to distance from the features. Given a choice between locations, an animal would invariably drift towards ones that are closer. Additionally, they also command some influence for longer distances. We did consider alternatives such as a potential that corresponds to an inverse squared force but it diminishes much faster as the distance to the source increases. The ‘repulsive features’ such as human dominated areas are incorporated using Gaussian type potentials that would have an influence only when the animal is close to them. Such forces fall off exponentially fast as one goes away from the source location.
    The corresponding Langevin equations can be written as:

    $$begin{aligned} frac{dx}{dt}= & {} -gamma sum _{i}frac{ 2alpha times (x – x_{i})}{(x – x_{i})^2 + (y – y_{i})^2} nonumber \&+,gamma sum _{j}Big ((x-x_{j}) e^{-((x-x_{j})^2 + (y-y_{j})^2)}Big ) + root 2 of {2D}xi _x(t) end{aligned}$$
    (1)

    $$begin{aligned} frac{dy}{dt}= & {} -gamma sum _{i}frac{ 2alpha times (y – y_{i})}{(x – x_{i})^2 + (y – y_{i})^2} nonumber \&+,gamma sum _{j}Big ((y-y_{j}) e^{-((x-x_{j})^2 + (y-y_{j})^2)}Big ) + root 2 of {2D}xi _y(t) end{aligned}$$
    (2)

    where x and y denote the coordinates of an animal’s location. ((x_{i}), (y_{i})) and ((x_{j}), (y_{j})) denote locations of i attracting and j repelling features respectively. We only choose nests as points of attraction for breeding hornbills since their diurnal movements are strongly centred around the nests. The white noise terms (xi _x) and (xi _y) are Gaussian in nature and delta correlated—which means that no correlations exist between the noise values at different instances of time. (gamma ) and D denote the drift and diffusion coefficients respectively. The drift coefficient (gamma ) represents the directedness of motion, which could be interpreted as strength of bias towards/against certain features in the landscape. In contrast, D quantifies the strength of random undirected motion. The force term with coefficient (-gamma ) results from negative gradient of the logarithmic potential, whose choice we explained earlier:

    $$begin{aligned} U = gamma sum _{i} log left{ (x – x_{i})^2 + (y – y_{i})^2 right} ^{alpha } ,,,. end{aligned}$$
    (3)

    The value of (alpha ) is determined from calculation of first passage times of the birds (discussed in the following section) and comparison of the values so obtained with observational (telemetry) data. We find that (alpha ) = 8 gives biologically sensible first passage times for hornbills (see “Calculating First Passage Times” in Methods section, Table 3 and Supplementary Tables 1, 2). If one observes an animal’s movement for a very long time, the probability of finding the animal would decrease more drastically away from a central feature for lower values of (alpha ). Such variations are captured by the steady-state probability distributions of space-use that we describe in the following section.
    Fokker–Planck methods
    Although the Langevin equations can generate trajectories of movement, the corresponding simulations need to be run for very long times to infer reliable information about spatial use. The time steps are further much smaller than the frequency of data recorded by the GPS. The step-lengths thus generated from simulated trajectories do not lend themselves to comparison against those from the recorded data. A convenient alternative is to solve a Fokker–Planck equation which has a direct correspondence with the Langevin equations. For our model, this takes the form:

    $$begin{aligned} frac{ partial P(x,y,t)}{partial t}&= frac{partial }{partial x} left{ F_x + D frac{partial }{partial x} right} P(x,y,t) nonumber \&quad +, frac{partial }{partial y} left{ F_y + D frac{partial }{partial y}.right} P(x,y,t) end{aligned}$$
    (4)

    where

    $$begin{aligned} F_x&= -gamma sum _{i} frac{ 2 alpha times (x – x_{i})}{(x – x_{i})^2 + (y – y_{i})^2} nonumber \&quad+, gamma sum _{j} (x – x_{j}) times e^{-( (x – x_{j})^2 + (y – y_{j})^2)} nonumber \ F_y&= -gamma sum _{i} frac{ 2 alpha times (y – y_{i})}{(x – x_{i})^2 + (y – y_{i})^2} nonumber \&quad+ gamma sum _{j} (y – y_{j}) times e^{-( (x – x_{j})^2 + (y – y_{j})^2)} end{aligned}$$
    (5)

    The Fokker–Planck equation describes the evolution of the probabilities of occurrence over a given region. The probability distribution eventually reaches a ‘steady state’ which captures the long-term occurrence probabilities for a given bird, and it does not change beyond this point in time. This steady-state probability distribution can be computed by setting the time derivative term to zero in Eq. (4). The numerical solution of the Fokker–Planck equation involves discretizing the spatial derivatives involved. The steady state probability distribution is consequently obtained on a spatial domain of discretized grids.
    Interestingly, Giuggioli et al.9 considered logarithmic potentials in their work on home range estimation, where an exponent of 8 was found to have a very similar steady state distribution to that from a harmonic potential. Harmonic potential has been utilized in analyzing home ranges of Peromyscus maniculatus10.
    Using the steady-state solution of the Fokker–Planck equation, we compute the mean square displacement averaged over different possible starting locations using the steady state distribution. A discrete version of the mean-square displacement (MSD) can be defined as:

    $$begin{aligned} MSD = sum _i^N langle (x – x_i)^2 + (y – y_i)^2 rangle P_{0}(x_i,y_i) end{aligned}$$
    (6)

    where (P_0(x_i,y_i)) is the distribution of starting locations (x_i) and (y_i) from where displacements are calculated. The inner angular brackets represent a similar weighted average of the mid-points of all grids over the steady-state probability distribution (P_{text {st}}(x,y)). Many of the grids that we define to perform simulations lie outside the known home range of the birds. The probability of choosing a starting location is defined using a Gaussian distribution centred around the nest or the most visited roost site.
    The square root of the MSD defines a characteristic length scale. This could be interpreted as home range length when the steady state distribution is computed over an infinite extent9. A logarithmic potential does not lend itself to such computations since it decays much more slowly such that the characteristic length continues to grow with the size of the area considered. We evaluate the characteristic length scale (L) on a domain that is not much larger in size compared to the observed home range.
    We also calculate L from empirical data by using the probability of occurrence over space inferred from two-dimensional histograms of location data. The MSD in this case is evaluated in the same vein as above but now the displacements from initial locations are weighted over the probabilities of occurrence derived from the histograms. Since these probabilities are only available for each grid, we choose only the mid-points of grids as possible locations to find the result. The starting locations are chosen from a uniform distribution over the mid-points of the grids. This is definitely a crude way of evaluating L but it does give us some way of comparing our numerical solutions against data. Finding a joint-probability distribution over the two dimensions would have been ideal but it is complicated by the fact that the distribution over space is multi-modal owing to multiple roosts for some hornbills. When inferring MSD from the location coordinates directly, it increases before saturating as the sampling frequency is decreased. For very high sampling frequency (or very small time intervals), diffusion effects dominate which leads to an almost linear increase in MSD. The effects of drift are more prominent compared to diffusion for lower sampling frequencies which marks the saturation of the MSD values10.
    A first-passage time model for heterogeneous environments
    The temporal information about an animal’s whereabouts is highly scrambled in the data. An important quantity of interest that could be extracted from movement data is the search time to reach a given target. A very useful measure of search times is the ‘first passage time’. Very generally, first passage time is the time taken for a given state variable to reach a particular value. In the case of animal movement, it can be interpreted as the time taken to reach a particular target location. McKenzie et al.11 derived an interesting first passage time model which had a direct correspondence with a Fokker–Planck equation. We use the prescription of Moorcroft et al.12,13 to estimate the drift and diffusion coefficients. This assumes a movement kernel that is a product of exponential distribution of step lengths and von Mises distribution for the turning angles. (This may be seen in the “goodness of fit tests” section in Methods where we assess fit of our data to claimed distributions.) It can be expressed as:

    $$begin{aligned} K({mathbf{X}} ,{mathbf{X}}’ ,tau )=, & {} frac{1}{rho } f_tau (rho ) k_tau (phi ) end{aligned}$$
    (7)

    $$begin{aligned} {rm{where}},,,,,,,,,,,, ,f_tau (rho )=, & {} lambda e^{-lambda rho }end{aligned}$$
    (8)

    $$begin{aligned} k_tau (phi )=, & {} frac{1}{2 pi I_0(kappa _tau )} exp [kappa _tau cos (phi )] end{aligned}$$
    (9)

    Here, ({mathbf{X}} ), ({mathbf{X}}’ ) denote the current and previous locations respectively, f is the exponential distribution of step lengths (rho ) with rate parameter (lambda ) and mean (bar{rho }_{tau } = 1/lambda ), and (k_{tau }) is the von Mises distribution of turning angles (phi ). (tau ) refers to the time taken to complete a given step. The turning angles are computed with respect to the nest/roost sites. (kappa _tau ) is the concentration parameter of the von Mises distribution which signifies the departure from a uniform distribution of movement directions. The normalizing factor (I_0(kappa _tau )) is a modified Bessel function of the first kind and of zeroth order. The drift and diffusion coefficients can be reliably estimated as:

    $$begin{aligned} gamma= & {} lim _{tau rightarrow 0} frac{bar{rho }_{tau } kappa _tau }{2tau } end{aligned}$$
    (10)

    $$begin{aligned} D= & {} lim _{tau rightarrow 0} frac{bar{{rho _{tau }}^2}}{4tau } end{aligned}$$
    (11)

    Employing the formalism in McKenzie11 to derive the equation for the first passage time T, we obtain the following equation:

    $$begin{aligned}&gamma sum _{i} left{ frac{ 2alpha times ({mathbf{X}} – {mathbf{X}} _{i})}{(x – x_{i})^2 + (y – y_{i})^2} right} cdot nabla T nonumber \&quad -, gamma sum _{j} left{ ({mathbf{X}} – {mathbf{X}} _{j}) e^{-( (x – x_{j})^2 + (y – y_{j})^2)} right} cdot nabla T nonumber \&quad +, D nabla ^2 T + 1 = 0 end{aligned}$$
    (12)

    The terms in dot product with (nabla T) are simply the drift coefficients with spatial dependence.
    McKenzie et al.11 had a simpler version of the first passage time equation that only accounted for bias towards the home range centre. The authors mention that the task of solving the first passage time equation is computationally harder with terms that account for more complex types of heterogeneities. We transform the partial differential equation in (12) into polar coordinates which simplifies the process of solving it (see First Passage Time calculation in Methods). The first passage times obtained from this solution also help us fix the value of (alpha ) in the equation above and subsequently in the logarithmic potential in (3), and in Eqs. (1) and (2). On performing this analysis for different hornbills, we see that (alpha ) = 8 works very well for them irrespective of the species and distribution of heterogeneities around them (see First Passage Time calculation in Methods). First passage times are calculated from the roosting/nesting site that lies closest to the home range centre. In case of GHNBr2, we calculate the first passage times from the approximate home range centre where no roosts exist. This ensures that most points considered for computations lie within the actual extent of the bird’s recorded locations. We used the Minimum Convex Polygon method to estimate the approximate home range centre14. This helped in identifying a location for each bird—which was a roost/nest in most cases—from where first passage times were subsequently computed. The method used for home range estimation is not relevant in the context of our proposed model and results presented, and therefore we do not consider other alternatives. More

  • in

    Effects of natural and experimental drought on soil fungi and biogeochemistry in an Amazon rain forest

    1.
    Lovejoy, T. E. & Nobre, C. Amazon tipping point. Sci. Adv. 4, eaat2340 (2018).
    Article  Google Scholar 
    2.
    Chadwick, R., Good, P., Martin, G. & Rowell, D. P. Large rainfall changes consistently projected over substantial areas of tropical land. Nat. Clim. Change 6, 177 (2015).
    Article  Google Scholar 

    3.
    Neelin, J. D., Münnich, M., Su, H., Meyerson, J. E. & Holloway, C. E. Tropical drying trends in global warming models and observations. Proc. Natl Acad. Sci. USA 103, 6110–6115 (2006).
    CAS  Article  Google Scholar 

    4.
    Barkhordarian, A., Saatchi, S. S., Behrangi, A., Loikith, P. C. & Mechoso, C. R. A recent systematic increase in vapor pressure deficit over tropical South America. Sci. Rep. 9, 15331 (2019).
    Article  CAS  Google Scholar 

    5.
    Cox, P. M. et al. Amazonian forest dieback under climate-carbon cycle projections for the 21st century. Theor. Appl. Climatol. 78, 137–156 (2004).
    Article  Google Scholar 

    6.
    Salazar, L. F., Nobre, C. A. & Oyama, M. D. Climate change consequences on the biome distribution in tropical South America. Geophys. Res. Lett. 34, L09708 (2007).
    Article  Google Scholar 

    7.
    Boisier, J. P., Ciais, P., Ducharne, A. & Guimberteau, M. Projected strengthening of Amazonian dry season by constrained climate model simulations. Nat. Clim. Change 5, 656 (2015).
    Article  Google Scholar 

    8.
    Amundson, R. & Jenny, H. On a state factor model of ecosystems. BioScience 47, 536–543 (1997).
    Article  Google Scholar 

    9.
    Schlesinger, W. H. et al. Forest biogeochemistry in response to drought. Glob. Change Biol. 22, 2318–2328 (2016).
    Article  Google Scholar 

    10.
    Bennett, E. M., Peterson, G. D. & Levitt, E. A. Looking to the future of ecosystem services. Ecosystems 8, 125–132 (2005).
    Article  Google Scholar 

    11.
    Phillips, O. L. et al. Drought sensitivity of the Amazon rain forest. Science 323, 1344–1347 (2009).
    CAS  Article  Google Scholar 

    12.
    Esquivel-Muelbert, A. et al. Compositional response of Amazon forests to climate change. Glob. Change Biol. 25, 39–56 (2019).
    Article  Google Scholar 

    13.
    Cox, P. M. et al. Sensitivity of tropical carbon to climate change constrained by carbon dioxide variability. Nature 494, 341 (2013).
    CAS  Article  Google Scholar 

    14.
    Eller, C. B. et al. Modelling tropical forest responses to drought and El Niño with a stomatal optimization model based on xylem hydraulics. Philos. Trans. R. Soc. B Biol. Sci. 373, 20170315 (2018).
    Article  CAS  Google Scholar 

    15.
    Meir, P. et al. Threshold responses to soil moisture deficit by trees and soil in tropical rain forests: insights from field experiments. Bioscience 65, 882–892 (2015).
    Article  Google Scholar 

    16.
    Davidson, E. A., Nepstad, D. C., Ishida, F. Y. & Brando, P. M. Effects of an experimental drought and recovery on soil emissions of carbon dioxide, methane, nitrous oxide, and nitric oxide in a moist tropical forest. Glob. Change Biol. 14, 2582–2590 (2008).
    Article  Google Scholar 

    17.
    da Costa, A. C. L. et al. Ecosystem respiration and net primary productivity after 8–10 years of experimental through-fall reduction in an eastern Amazon forest. Plant Ecol. Diversity 7, 7–24 (2014).
    Article  Google Scholar 

    18.
    Doughty, C. E. et al. Drought impact on forest carbon dynamics and fluxes in Amazonia. Nature 519, 78 (2015).
    CAS  Article  Google Scholar 

    19.
    Fisher, R. A., Williams, M., Do Vale, R. L., Da Costa, A. L. & Meir, P. Evidence from Amazonian forests is consistent with isohydric control of leaf water potential. Plant Cell Environ. 29, 151–165 (2006).
    Article  Google Scholar 

    20.
    Rowland, L. et al. Death from drought in tropical forests is triggered by hydraulics not carbon starvation. Nature 528, 119–122 (2015).
    CAS  Article  Google Scholar 

    21.
    Nepstad, D. C., Tohver, I. M., Ray, D., Moutinho, P. & Cardinot, G. Mortality of large trees and lianas following experimental drought in a Amazon forest. Ecology 88, 2259–2269 (2007).
    Article  Google Scholar 

    22.
    da Costa, A. C. L. et al. Effect of 7 yr of experimental drought on vegetation dynamics and biomass storage of an eastern Amazonian rain forest. N. Phytol. 187, 579–591 (2010).
    Article  Google Scholar 

    23.
    Rowland, L. et al. Shock and stabilisation following long-term drought in tropical forest from 15 years of litterfall dynamics. J. Ecol. 106, 1673–1682 (2018).
    Article  Google Scholar 

    24.
    Sotta, E. D. et al. Effects of an induced drought on soil carbon dioxide (CO2) efflux and soil CO2 production in an Eastern Amazonian rain forest, Brazil. Glob. Change Biol. 13, 2218–2229 (2007).
    Article  Google Scholar 

    25.
    Bardgett, R. D. & van der Putten, W. H. Belowground biodiversity and ecosystem functioning. Nature 515, 505–511 (2014).
    CAS  Article  Google Scholar 

    26.
    Koyama, A., Steinweg, J. M., Haddix, M. L., Dukes, J. S. & Wallenstein, M. D. Soil bacterial community responses to altered precipitation and temperature regimes in an old field grassland are mediated by plants. FEMS Microbiol. Ecol. https://doi.org/10.1093/femsec/fix156 (2017).

    27.
    Kivlin, S. N. & Hawkes, C. V. Tree species, spatial heterogeneity, and seasonality drive soil fungal abundance, richness, and composition in Neotropical rain forests. Environ. Microbiol. 18, 4662–4673 (2016).
    Article  Google Scholar 

    28.
    Sinsabaugh, R. L. & Moorhead, D. L. Resource allocation to extracellular enzyme production: a model for nitrogen and phosphorus control of litter decomposition. Soil Biol. Biochem. 26, 1305–1311 (1994).
    Article  Google Scholar 

    29.
    Turner, B. L. & Romero, T. E. Stability of hydrolytic enzyme activity and microbial phosphorus during storage of tropical rain forest soils. Soil Biol. Biochem. 42, 459–465 (2010).
    CAS  Article  Google Scholar 

    30.
    Sinsabaugh, R. L. et al. Stoichiometry of soil enzyme activity at global scale. Ecol. Lett. 11, 1252–1264 (2008).
    Article  Google Scholar 

    31.
    Waring, B. G., Weintraub, S. R. & Sinsabaugh, R. L. Ecoenzymatic stoichiometry of microbial nutrient acquisition in tropical soils. Biogeochemistry 117, 101–113 (2014).
    CAS  Article  Google Scholar 

    32.
    Turner, B. L. & Joseph Wright, S. The response of microbial biomass and hydrolytic enzymes to a decade of nitrogen, phosphorus, and potassium addition in a lowland tropical rain forest. Biogeochemistry 117, 115–130 (2014).
    CAS  Article  Google Scholar 

    33.
    Weintraub, S. R., Wieder, W. R., Cleveland, C. C. & Townsend, A. R. Organic matter inputs shift soil enzyme activity and allocation patterns in a wet tropical forest. Biogeochemistry 114, 313–326 (2013).
    CAS  Article  Google Scholar 

    34.
    Firestone, M. K. & Davidson, E. A. in Exchange of Trace Gases between Terrestrial Ecosystems and The Atmosphere (eds. M. O. Andreae & D. S. Schimel) 7–21 (John Wiley and Sons, 1989).

    35.
    Meir, P. et al. Short-term effects of drought on tropical forest do not fully predict impacts of repeated or long-term drought: gas exchange versus growth. Philos. Trans. R. Soc. B Biol. Sci. 373, 20170311 (2018).
    Article  Google Scholar 

    36.
    Robertson, G. P. in Mineral Nutrients in Troical Forest and Savanna Ecosystems (ed. J. Proctor) 55–69 (Blackwell Scientific, 1989).

    37.
    Silver, W. L., Lugo, A. E. & Keller, M. Soil oxygen availability and biogeochemistry along rainfall and topographic gradients in upland wet tropical forest soils. Biogeochemistry 44, 301–328 (1999).
    Google Scholar 

    38.
    Pett-Ridge, J. & Firestone, M. K. Redox fluctuation structures microbial communities in a wet tropical soil. Appl. Environ. Microbiol. 71, 6998–7007 (2005).
    CAS  Article  Google Scholar 

    39.
    Cleveland, C. C., Reed, S. C. & Townsend, A. R. Nutrient regulation of organic matter decomposition in a tropical rain forest. Ecology 87, 492–503 (2006).
    Article  Google Scholar 

    40.
    Cleveland, C. C., Wieder, W. R., Reed, S. C. & Townsend, A. R. Experimental drought in a tropical rain forest increases soil carbon dioxide losses to the atmosphere. Ecology 91, 2313–2323 (2010).
    Article  Google Scholar 

    41.
    Wallenstein, M. D. & Hall, E. K. A trait-based framework for predicting when and where microbial adaptation to climate change will affect ecosystem functioning. Biogeochemistry 109, 35–47 (2012).
    Article  Google Scholar 

    42.
    Knapp, A. K. et al. Consequences of more extreme precipitation regimes for terrestrial ecosystems. BioScience 58, 811–821 (2008).
    Article  Google Scholar 

    43.
    Greenland, D., Goodin, D. G. & Smith, R. C. Climate Variability and Ecosystem Response at Long-Term Ecological Research Sites. (Oxford University Press, 2003).

    44.
    Kayler, Z. E. et al. Experiments to confront the environmental extremes of climate change. Science 13, 219–225 (2015).
    Google Scholar 

    45.
    McGuire, K. L., Fierer, N., Bateman, C., Treseder, K. K. & Turner, B. L. Fungal community composition in neotropical rain forests: the influence of tree diversity and precipitation. Microb. Ecol. 63, 804–812 (2012).
    Article  Google Scholar 

    46.
    Buscardo, E. et al. Spatio-temporal dynamics of soil bacterial communities as a function of Amazon forest phenology. Sci. Rep. 8, 4382 (2018).
    Article  CAS  Google Scholar 

    47.
    Quesada, C. A. et al. Variations in chemical and physical properties of Amazon forest soils in relation to their genesis. Biogeosciences 7, 1515–1541 (2010).
    CAS  Article  Google Scholar 

    48.
    Bonfim, J. A., Vasconcellos, R. L. F., Baldesin, L. F., Sieber, T. N. & Cardoso, E. Dark septate endophytic fungi of native plants along an altitudinal gradient in the Brazilian Atlantic forest. Fung. Ecol. 20, 202–210 (2016).
    Article  Google Scholar 

    49.
    Carson, J. K. et al. Low pore connectivity increases bacterial diversity in soil. Appl. Environ. Microbiol. 76, 3936–3942 (2010).
    CAS  Article  Google Scholar 

    50.
    Schimel, J. & Schaeffer, S. Microbial control over carbon cycling in soil. Front. Microbiol. 3, 1–11 (2012).
    Article  CAS  Google Scholar 

    51.
    Daws, S. C. et al. Do shared traits create the same fates? Examining the link between morphological type and the biogeography of fungal and bacterial communities. Fung. Ecol. 46, 100948 (2020).
    Article  Google Scholar 

    52.
    DeAngelis, K. M., Silver, W. L., Thompson, A. W. & Firestone, M. K. Microbial communities acclimate to recurring changes in soil redox potential status. Environ. Microbiol. 12, 3137–3149 (2010).
    CAS  Article  Google Scholar 

    53.
    Bradford, M. A. et al. Thermal adaptation of soil microbial respiration to elevated temperature. Ecol. Lett. 11, 1316–1327 (2008).
    Article  Google Scholar 

    54.
    Coleman, D. C., Callaham, M. A. Jr. & Crossley, D. A. Jr. Fundamentals of Soil Ecology 3rd edn. (Academic Press, 2018).

    55.
    de Meester, L. in Biogeography of Microscopic Organisms: Is Everything Small Everywhere? (ed. D. Fontaneto) 324–334 (Cambridge University Press, 2011).

    56.
    Leibold, M. A. et al. The metacommunity concept: a framework for multi-scale community ecology. Ecology Letters 7, 601–613 (2004).
    Article  Google Scholar 

    57.
    Barberán, A. et al. Continental-scale distributions of dust-associated bacteria and fungi. Proc. Natl Acad. Sci. USA 112, 5756–5761 (2015).
    Article  CAS  Google Scholar 

    58.
    Cáliz, J., Triadó-Margarit, X., Camarero, L. & Casamayor, E. O. A long-term survey unveils strong seasonal patterns in the airborne microbiome coupled to general and regional atmospheric circulations. Proc. Natl Acad. Sci. USA 115, 12229–12234 (2018).
    Article  CAS  Google Scholar 

    59.
    Prospero, J. M., Glaccum, R. A. & Nees, R. T. Atmospheric transport of soil dust from Africa to South America. Nature 289, 570–572 (1981).
    CAS  Article  Google Scholar 

    60.
    Rime, T., Hartmann, M. & Frey, B. Potential sources of microbial colonizers in an initial soil ecosystem after retreat of an alpine glacier. ISME J. 10, 1625 (2016).
    CAS  Article  Google Scholar 

    61.
    Elena, S. F. & Lenski, R. E. Evolution experiments with microorganisms: the dynamics and genetic bases of adaptation. Nat. Rev. Genet. 4, 457–469 (2003).
    CAS  Article  Google Scholar 

    62.
    terHorst, C. P., Lennon, J. T. & Lau, J. A. The relative importance of rapid evolution for plant-microbe interactions depends on ecological context. Proc. R. Soc. B Biol. Sci. 281, 20140028 (2014).
    Article  Google Scholar 

    63.
    Read, D. J. & Haselwandter, K. Observations on the mycorrhizal status of some alpine plant communities. N. Phytol. 88, 341–352 (1981).
    Article  Google Scholar 

    64.
    Bell, A. A. & Wheeler, M. H. Biosynthesis and functions of fungal melanins. Ann. Rev. Phytopathol. 24, 411–451 (1986).
    CAS  Article  Google Scholar 

    65.
    Mandyam, K. & Jumpponen, A. Seeking the elusive function of the root-colonising dark septate endophytic fungi. Studies Mycol. 53, 173–189 (2005).
    Article  Google Scholar 

    66.
    da Costa, A. C. L. et al. Stand dynamics modulate water cycling and mortality risk in droughted tropical forest. Glob. Change Biol. 24, 249–258 (2018).
    Article  Google Scholar 

    67.
    Jumpponen, A. & Trappe, J. M. Dark septate endophytes: a review of facultative biotrophic root-colonizing fungi. N. Phytol. 140, 295–310 (1998).
    Article  Google Scholar 

    68.
    Fisher, R. A. et al. The response of an Eastern Amazonian rain forest to drought stress: results and modelling analyses from a throughfall exclusion experiment. Glob. Change Biol. 13, 2361–2378 (2007).
    Article  Google Scholar 

    69.
    Newsham, K. K. A meta-analysis of plant responses to dark septate root endophytes. N. Phytol. 190, 783–793 (2011).
    CAS  Article  Google Scholar 

    70.
    Smith, S. E. & Read, D. J. Mycorrhizal symbiosis. 3rd edn. (Academic Press, 2008).

    71.
    IPCC. Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. (Cambridge University Press, 2013).

    72.
    Janzen, D. H. Herbivores and the number of tree species in tropical forests. Am. Natural. 104, 501–528 (1970).
    Article  Google Scholar 

    73.
    Connell, J. H. in Dynamics of Populations (eds. P. J. den Boer & G. R. Gradwell) 298–312 (Centre for Agricultural Publishing and Documentation, 1971).

    74.
    Comita, L. S. et al. Testing predictions of the Janzen–Connell hypothesis: a meta-analysis of experimental evidence for distance- and density-dependent seed and seedling survival. J. Ecol. 102, 845–856 (2014).
    Article  Google Scholar 

    75.
    Buscardo, E. et al. in Interactions between Biosphere, Atmosphere and Human Land Use in The Amazon Basin (eds. Laszlo Nagy, Bruce R. Forsberg, & Paulo Artaxo) 225–266 (Springer Berlin Heidelberg, 2016).

    76.
    Singh, J. S., Raghubanshi, A. S., Singh, R. S. & Srivastava, S. C. Microbial biomass acts as a source of plant nutrients in dry tropical forest and savanna. Nature 338, 499 (1989).
    Article  Google Scholar 

    77.
    Luizão, F., Luizão, R. & Chauvel, A. Premiers résultats sur la dynamique des biomasses racinaires et microbiennes dans un latosol d’Amazonie centrale (Brésil) sous forêt et sous pâturage. Cahiers ORSTOM. Série Pédologie 27, 69–79 (1992).
    Google Scholar 

    78.
    Vasconcelos, H. L. & Luizão, F. J. Litter production and litter nutrient concentrations in a fragmented Amazonian landscape. Ecol. Appl. 14, 884–892 (2004).
    Article  Google Scholar 

    79.
    Cornejo, F. H., Varela, A. & Wright, S. J. Tropical forest litter decomposition under seasonal drought: nutrient release, fungi and bacteria. Oikos 70, 183–190 (1994).
    Article  Google Scholar 

    80.
    Garcia-Montiel, D. C. et al. Controls on soil nitrogen oxide emissions from forest and pastures in the Brazilian Amazon. Glob. Biogeochem. Cycles 15, 1021–1030 (2001).
    CAS  Article  Google Scholar 

    81.
    Malhi, Y. et al. The productivity, metabolism and carbon cycle of two lowland tropical forest plots in south-western Amazonia, Peru. Plant Ecol. Diversity 7, 85–105 (2014).
    Article  Google Scholar 

    82.
    Cleveland, C. C. & Townsend, A. R. Nutrient additions to a tropical rain forest drive substantial soil carbon dioxide losses to the atmosphere. Proc. Natl Acad. Sci. USA 103, 10316–10321 (2006).
    CAS  Article  Google Scholar 

    83.
    Allison, S. D., Weintraub, M. N., Gartner, T. B. & Waldrop, M. P. in Soil enzymology (eds. Girish Shukla & Ajit Varma) 229–243 (Springer Berlin Heidelberg, 2011).

    84.
    Classen, A. T. et al. Direct and indirect effects of climate change on soil microbial and soil microbial-plant interactions: What lies ahead? Ecosphere 6, 1–21 (2015).
    Article  Google Scholar 

    85.
    Hoeksema, J. D. et al. Ectomycorrhizal plant-fungal co-invasions as natural experiments for connecting plant and fungal traits to their ecosystem consequences. Front. Forests Glob. Change https://doi.org/10.3389/ffgc.2020.00084 (2020).

    86.
    Allison, S. D. Cheaters, diffusion and nutrients constrain decomposition by microbial enzymes in spatially structured environments. Ecol. Lett. 8, 626–635 (2005).
    Article  Google Scholar 

    87.
    Nottingham, A. T. et al. Microbes follow Humboldt: temperature drives plant and soil microbial diversity patterns from the Amazon to the Andes. Ecology 99, 2455–2466 (2018).
    Article  Google Scholar 

    88.
    Štursova, M., Crenshaw, C. L. & Sinsabaugh, R. L. Microbial responses to long-term N deposition in a semiarid grassland. Microb. Ecol. 51, 90–98 (2006).
    Article  Google Scholar 

    89.
    Henry, H. A. L., Juarez, J. D., Field, C. B. & Vitousek, P. M. Interactive effects of elevated CO2, N deposition and climate change on extracellular enzyme activity and soil density fractionation in a California annual grassland. Glob. Change Biol. 11, 1808–1815 (2005).
    Article  Google Scholar 

    90.
    Lashermes, G., Gainvors-Claisse, A., Recous, S. & Bertrand, I. Enzymatic strategies and carbon use efficiency of a litter-decomposing fungus grown on maize leaves, stems, and roots. Front. Microbiol. 7, 1315 (2016).

    91.
    Chet, I. in Innovative Approaches to Plant Disease Control (ed. I. Chet) (Wiley, 1987).

    92.
    Boller, T. in Cellular and Molecular Biology of Plant Stress (eds. J. L. Key & T. Kosuge) (Liss, A.R., 1985).

    93.
    Bond-Lamberty, B., Bailey, V. L., Chen, M., Gough, C. M. & Vargas, R. Globally rising soil heterotrophic respiration over recent decades. Nature 560, 80–83 (2018).
    CAS  Article  Google Scholar 

    94.
    Ruivo, M. & Cunha, E. in Ecosystems and Sustainable Development (eds. E. Tiezzi, C. A. Brebbia, & J. L. Uso) 1113–1121 (WIT Press, 2003).

    95.
    Eastman, J. R. TerrSet Manual (Clark University, 2015).

    96.
    Ihrmark, K. et al. New primers to amplify the fungal ITS2 region—evaluation by 454-sequencing of artificial and natural communities. FEMS Microbiol. Ecol. 82, 666–677 (2012).
    CAS  Article  Google Scholar 

    97.
    White, T. J., Bruns, T. D., Lee, S. B. & Taylor, J. W. in PCR—Protocols and applications—A laboratory manual (eds. N. Innis, D. Gelfand, J. Sninsky, & T. White) 315–322 (Academic Press, 1990).

    98.
    Lindahl, B. D. et al. Fungal community analysis by high-throughput sequencing of amplified markers—a user’s guide. N. Phytol. 199, 288–299 (2013).
    CAS  Article  Google Scholar 

    99.
    Kõljalg, U. et al. Towards a unified paradigm for sequence-based identification of fungi. Mol. Ecol. 22, 5271–5277 (2013).
    Article  CAS  Google Scholar 

    100.
    Nguyen, N. H. et al. FUNGuild: an open annotation tool for parsing fungal community datasets by ecological guild. Fung. Ecol. 20, 241–248 (2016).
    Article  Google Scholar 

    101.
    Oksanen, J. et al. vegan: community ecology package. R package version 2.3-5. http://CRAN.R-project.org/package=vegan. (2016).

    102.
    Pritsch, K. et al. Optimized assay and storage conditions for enzyme activity profiling of ectomycorrhizae. Mycorrhiza 21, 589–600 (2011).
    CAS  Article  Google Scholar 

    103.
    Souza, R. C. et al. Responses of soil extracellular enzyme activities to experimental warming and CO2 enrichment at the alpine treeline. Plant Soil 416, 527–537 (2017).
    CAS  Article  Google Scholar 

    104.
    Baselga, A. & Orme, C. D. L. betapart: an R package for the study of beta diversity. Methods Ecol. Evol. 3, 808–812 (2012).
    Article  Google Scholar 

    105.
    Vohník, M. & Albrechtová, J. The co-occurrence and morphological continuum between ericoid mycorrhiza and dark septate endophytes in roots of six european Rhododendron species. Folia Geobotanica 46, 373–386 (2011).
    Article  Google Scholar 

    106.
    Grelet, G., Martino, E., Dickie, I. A., Tajuddin, R. & Artz, R. in Molecular mycorrhizal Symbiosis (ed. F. Martin) (John Wiley & Sons, Inc, 2017).

    107.
    Benjamini, Y. & Hochberg, Y. Controlling the false diiscovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B 57, 289–300 (1995).
    Google Scholar  More

  • in

    How to limit the ecological costs of urbanization in China

    Linjun Xie is a postdoctoral researcher studying urban sustainability and environmental governance at Durham University, UK.Credit: Samer Angelone

    Postdoctoral researcher Linjun Xie reveals what an eco-island on the outskirts of Shanghai taught her about sustainable development in China
    What is your research area?
    I study urban sustainability and environmental governance at Durham University, UK, although I’ve been home in China throughout the pandemic.
    In recent decades, rural areas close to megacities such as Shanghai, Beijing and Shenzhen have been absorbed into city development plans. Chongming is one such rural area, made up of three islands in the mouth of the Yangtze River, northeast of Shanghai. The Chongming Eco-Island Project is a municipal government scheme intended to be a model for more environmentally sustainable urban development in China.
    These urban transitions can alter local landscapes and ecology, causing a loss of wildlife and natural habitats, as well as environmental pollution.
    So in 2010, Shanghai announced its ambition to turn the Chongming district into a modern eco-island that would balance ecological sustainability with economic growth.
    How did you end up researching the impact of sustainable development in China from the United Kingdom?
    After completing my degree in urban planning at Huaqiao University in Xiamen, China, I was looking for places to research sustainable development. Cardiff University in the United Kingdom offers a year-long master’s course as part of its eco-cities research programme and I joined in 2015. During my course, I heard many references to how Chongming was different from other eco-projects and so when I applied to do my PhD at the University of Nottingham Ningbo in China, I asked if I could focus on it for my thesis.
    What makes Chongming different from other sustainable development projects?
    Chongming is unlike other state-led eco-projects in China, such as the Tianjin eco-city, a collaborative effort between Singapore and China, or the Shenzhen International Low Carbon City. These highly compact and modern cities are constructed on empty land or previous industrial sites. In these cities, everyone is a newcomer: there are no indigenous people.

    The district of Chongming, by contrast, is an environment with high-quality rural land, diverse landscapes ranging from wetlands and crop fields to forests, and it is already home to nearly 700,000 people. Their homes are spread over a large area so it feels sparsely populated. Many people have lived on the islands their whole lives. This means that you can’t start from scratch. Policies need to be integrated into the local community and ecology.
    You might wonder why plans to turn a rural community into an eco-project are necessary at all. The answer is that without protections, the area will not stay this way for long. The land is close to central Shanghai and so has development value. One part of the area is not part of this eco-project and you can see how quickly high-rise buildings have gone up.
    In 2010, a list of goals were set by the Chongming district government, including limits on construction, the protection of arable land and an increase in forestation.
    It also set social and economic targets, such as the implementation of clean-energy transportation, consolidation of the population into compact settlements, and development of the island’s green industries, such as organic farming and low-carbon manufacturing.
    How successful is this project?
    Statistics show that unbridled urbanization, which is common in China, has been reversed in Chongming. For instance, by 2016 the forest coverage on Chongming was 23%, twice the average of Shanghai.
    A series of ecological tourism projects have been built, such as the Dongtan Wetland Park, and tourism revenue quadrupled from 2008 to 2016, rising from 270 million yuan (US$41.7 million) to 1.09 billion yuan.
    In what areas could the project be improved?
    Our research revealed some concerns. For example, the targets set in the eco-island plan also serve as key evaluation criteria for officials’ job performance. So they encourage the adoption of short-term measures that are not necessarily the best long-term solutions.
    For example, to increase the amount of forest cover, extensive land has been turned into forest, but plantations of a single fast-growing tree species have been introduced that do not encourage or support local biodiversity.

    Also, the aesthetics of the landscape are sometimes prioritized over the needs of local ecology and biodiversity: cement is often used, and uniformly landscaped riverbanks for river regulation are common. These are an attempt to improve the water quality in rivers but don’t support local wetland plants and aquatic species.
    There is also the question of transport. Chongming is an attractive rural retreat for Shanghai residents and on weekends and during national holidays, the Shanghai Yangtze River Tunnel-Bridge, which directly connects the east of Chongming Island to central Shanghai, is often terribly congested.
    Public transport needs to improve. Once you arrive on the island, there are electric buses and many tourists use bicycles, but cars are still more convenient when making the journey from Shanghai.
    Chongming must find a balance between its economic and ecological interests. The region needs the money that comes from tourism. But to be truly sustainable it needs to become both self-reliant and environmentally secure.
    In general, what needs to be done to achieve sustainable urban development?
    Well-intentioned ecological initiatives can, in fact, have destructive effects if the locality is not completely understood. For example, in Chongming, government officials adopted a strategy called ‘one town, one tree species and one flower’, which meant that each local town needed to plant a different tree species and flower species in their respective jurisdiction.
    The selection of plants was chosen at random from a list produced by the Chongming district government. Consequently, more than 20 species of trees were planted separately in each town, including imported varieties such as northern red oak (Quercus rubra) and red maple (Acer rubrum).
    But this plan poses risks for biodiversity because new plants can destroy the natural connectivity between local species.
    So it is crucial to foster connections between historians, ecologists, engineers, planners, policymakers and local communities when planning and building ecological development. More

  • in

    Biodiversity’s importance is growing in China’s urban agenda

    Many cities in China, such as Xi’an (pictured), have experienced rapid growth in the past few decades.Credit: Xinhua/Shutterstock

    On 28 January 2020, a team of Chinese conservation scientists distributed a questionnaire across social-media platforms, asking Chinese citizens how they felt about proposed legislation that would ban the consumption and trade of wildlife in the country.
    It was an apposite moment: the questionnaire hit social-media platforms such as WeChat and Weibo just days after China had been forced to close its major cities to prevent the spread of a disease that scientists suspected was transferred to humans from an animal species at a market in Wuhan.
    More than 90% of the 74,070 respondents were in favour of a complete ban on wildlife trade — and, a month later, the central government came to the same conclusion and legislated to that effect. Researchers are increasingly studying the impact of these policies, and the country’s biodiversity. But big questions remain about whether China will deliver on its growing list of environmental commitments.

    Bin Zhao, an ecologist at Fudan University in Shanghai, China, says that, since the start of the COVID-19 pandemic, people in urban areas have been paying more attention to biodiversity than ever before. “People realized that contact with wild animals could lead to an outbreak of an epidemic, even in urban areas. This not only enhanced people’s understanding of biodiversity, but also promoted the idea that wildlife-protection law needed to be improved,” says Zhao.
    It came at a time when China was already committed to changing its approach to ecological protection, he says. In 2018, China amended its constitution to include the goal of becoming an ‘ecological civilization’. In the words of Chinese President Xi Jinping in 2017, economic development could no longer be at the expense of the environment.
    Multiple environmentally friendly policies have already been announced, such as the introduction of an ‘ecological red line’ policy to protect more of the Chinese mainland from development (see ‘Protected land’); a new network of national parks; stricter supervision of conservation; and a streamlining of environmental-oversight agencies — all to meet a government target of making the country’s environment ‘beautiful’ by 2035.

    Sources: UN/Xinhua/OECD

    Big cities, few controls
    In 1950, about only 13% of China’s population lived in cities. But since the 1980s, the country’s cities have grown rapidly as the engines of its economic growth (see ‘Urban population’). Millions left homes in rural areas to forge more prosperous lives in growing and newly built cities. Government policies, aimed at bolstering the economy, helped to encourage close to two-thirds of China’s population to move to these new urban areas, and the nation continues to have one of the world’s fastest growing urban populations. This has put intense pressure on the country’s ecology.

    Sources: UN/Xinhua/OECD

    “From an economic perspective, our ecosystems and environment have historically been considered to be worthless,” says Zhao. China’s natural resources, such as its wetlands, forests and water sources, haven’t received the same level of care from authorities as targets for economic growth, he says (see ‘Vegetation change’).

    Sources: UN/Xinhua/OECD

    As urban areas grow, there are direct and indirect impacts on ecological systems, according to Rob McDonald, who researches the impact and dependencies of cities on the natural world at The Nature Conservancy in Washington DC.
    Land is repurposed for development, and natural resources are needed to construct buildings and provide food and water for city dwellers, he says. These changes can lead to environmental problems, such as water and air pollution, insufficient water availability and deforestation much farther afield than in urban areas themselves.
    China’s government has been open about its commitment to tackling these problems, says Alice Hughes, a zoologist at the Xishuangbanna Tropical Botanical Garden in Menglun town, China. In May 2021, China will host the fifteenth United Nations Convention on Biological Diversity, also known as COP 15, in Kunming, where 200 countries will meet to sign off on a legally binding set of global targets to protect the world’s biodiversity. The country has already contributed to some broader environmental targets, including being carbon neutral by 2060.
    China has had some success, most notably in reducing air pollution. For example, in 2017, the amount of fine particulate matter in Beijing’s air dropped by just under 40% from the level in 2013, the year when national targets were launched.
    But at a press conference to discuss China’s progress on ecological and environmental protection, Cui Shuhong, an official at the Ministry of Ecology and Environment, said the country has much more to do to alleviate the fundamental pressures placed on its natural resources by economic development.
    Zhengguang Zhu, an environmental officer at China’s National Marine Environmental Monitoring Center, is familiar with preparations for COP 15: there are multiple working groups operating within the Ministry of Ecology and Environment, which are each responsible for different aspects of the event, from logistics to setting targets for improvements to China’s environment.

    Live turtles on display at a wildlife market in Shanghai, China, in August 2020. During the COVID-19 pandemic, the Chinese government issued a policy banning wildlife trade for food, but trade of exotic animals as pets still continues.Credit: Ales Plavevski/EPA-EFE/Shutterstock

    These working groups ask China’s public bodies, such as the ministry of agriculture, to offer their opinions on what the country feels should be included in the final roadmap for the coming decade.
    “I think the meeting will show that China has done its homework and is willing to be a good host. But leadership is not just about hospitality. It’s about having an ambitious framework that enables change, and I think we’ve got a long way to go before that happens,” says Zhu.
    Behaviour change
    Conservation researcher Tien Ming Lee, based at the Sun Yat-sen University in Guangzhou, China, says scientists and politicians are currently focused on finding better ways to protect Chinese ecosystems while continuing the country’s urban economic growth.
    His research team works across a range of projects, all focused on finding ways to prompt people to act differently and sustainably. For example, he is currently part of a 4-year, €10-million (US$12 million) project, mainly funded by the European Union, called Partners against Wildlife Crime. The project, which began in January 2019, hopes to disrupt the illicit supply chains through which exotic animals and plants, specifically tigers (Panthera tigris), Asian elephants (Elephas maximus), Siamese rosewood (Dalbergia cochinchinensis) and freshwater turtles, are traded throughout Cambodia, China, Laos, Malaysia, Myanmar, Thailand and Vietnam.
    As part of this project, Lee’s team and Lishu Li at the Wildlife Conservation Society China Counter Wildlife Trafficking Program are developing marketing materials to change the buying habits of urban Chinese consumers by attempting to dissuade them from illegal acts, such as buying tiger bone or elephant skin online for jewellery and traditional medicine, or keeping endangered freshwater turtles as pets. Lee says the materials have been developed with behavioural-science techniques: they aim to appeal to consumers’ desire to be seen to act in a conscientious manner.

    Police patrol the wetlands of the Yellow River Estuary ecotourism area near Dongying City, China.Credit: Costfoto/Barcroft Media via Getty

    Lee has also been part of a research project that looked at how trade agreements that stem from the country’s international Belt and Road economic initiative, an infrastructure project that aims to link trade across Europe, Asia and Africa to China, could lead to a greater demand for traditional Chinese medicine across the world. The plant, animal and fungal products used in these practises are often sourced from the wild, which might exacerbate the illegal and unsustainable trade of those species, he says.
    His research, a collaboration with Amy Hinsley, a conservation biologist at the University of Oxford, UK, concluded that there was a clear, urgent need for China to introduce carefully managed supply chains and ensure that rural farmers have resources for sustainable farming.
    During her four-decade career, Lu Zhi, a conservation biologist at Peking University in Beijing, has seen a shift in her field’s focus. It moved from observing animals in their natural habitats and coming up with ways to protect them from human activity to observing human behaviour: studying what can be done to make people’s lives more ecologically sustainable.
    In 2017, Zhi’s Shanshui Conservation Center, a non-governmental organization she founded in 2007 to develop community-based conservation projects, began working with herdsmen in Qinghai province on the Tibetan Plateau. The team wanted to help them to develop livelihoods from conservation activities in an underdeveloped, highly biodiverse area of China. The villagers learnt how to patrol and monitor wildlife, and how to act as guides for tourists interested in animal watching — including for the elusive and endangered snow leopard (Panthera uncia). Similar projects have been rolled out in 42 villages in western China.
    Zhi admits that such small projects are certainly not enough to bring the paradigm shift needed to safeguard the country’s vulnerable ecosystems. Government intervention has proved to be effective in tackling the larger issues, such as air and water pollution, she says. But “it’s not fair to ask people in rural areas not to develop their lives for the sake of wildlife, while others live in prosperous cities. We need alternative solutions.” More

  • in

    Malaria trends in Ethiopian highlands track the 2000 ‘slowdown’ in global warming

    1.
    Pascual, M., Ahumada, J., Chaves, L. F., Rodó, X. & Bouma, M. Malaria resurgence in East African Highlands: temperature trends revisited. Proc. Natl Acad. Sci. USA 103, 5829–5834 (2006).
    ADS  CAS  PubMed  Article  Google Scholar 
    2.
    Alonso, D., Bouma, M. J. & Pascual, M. Epidemic malaria and warmer temperatures in recent decades in an East African highland. Proc. Roy. Soc. B Biol. Sci. 278, 1661–1669 (2011).
    Google Scholar 

    3.
    Stern, D. I. et al. Temperature and malaria trends in highland East Africa. PLoS ONE 6, e24524 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    4.
    Chaves, L. F. & Koenraadt, C. J. M. Climate change and highland malaria: fresh air for a hot debate. Quart. Rev. Biol. 85, 27–55 (2010).
    PubMed  Article  Google Scholar 

    5.
    Shanks, G. D., Hay, S. I., Omumbo, J. A. & Snow, R. W. Malaria in Kenya’s Western Highlands. Emerg. Infect. Dis. 11, 1425–1432 (2005).
    PubMed  PubMed Central  Article  Google Scholar 

    6.
    Omumbo, J. A., Lyon, B., Waweru, S. M., Connor, S. J. & Thomson, M. C. Raised temperatures over the Kericho tea estates: revisiting the climate in the East African highlands malaria debate. Mal. J. 10, 12 (2011).
    Article  Google Scholar 

    7.
    Siraj, A. S. et al. Altitudinal changes in malaria incidence in highlands of Ethiopia and Colombia. Science 343, 1154–1158 (2014).
    ADS  CAS  PubMed  Article  Google Scholar 

    8.
    Caminade, C. et al. Climate change and malaria: model intercomparison. Proc. Natl Acad. Sci. USA 111, 3286–3291 (2014).

    9.
    Mordecai, E. A. et al. Thermal biology of mosquito-borne disease. Ecol. Lett. 22, 1690–1708 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    10.
    Shapiro, L. L. M., Whitehead, S. A. & Thomas, M. B. Quantifying the effects of temperature on mosquito and parasite traits that determine the transmission potential of human malaria. PLoS Biol. 15, e2003489 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    11.
    Waite, J. L., Suh, E., Lynch, P. A., & Thomas, M. B. Exploring the lower thermal limits for development of the human malaria parasite, Plasmodium falciparum. Biol. Lett. 15, 20190275 (2019).

    12.
    Brohan, P., Kennedy, J. J., Harris, I., Tett, S. F. & Jones, P. D. Uncertainty estimates in regional and global observed temperature changes: a new dataset from 1850. J. Geophys. Res. Atmos. 111, D12106 (2006).
    ADS  Article  Google Scholar 

    13.
    Kerr, R. What happened to global warming? Scientists say just wait a bit. Science 326, 28–29 (2009).
    ADS  CAS  PubMed  Article  Google Scholar 

    14.
    Meehl, G. A., Arblaster, J. M., Fasullo, J. T., Hu, A. & Trenberth, K. E. Model-based evidence of deep-ocean heat uptake during surface-temperature hiatus periods. Nat. Clim. Change 1, 360–364 (2011).
    ADS  Article  Google Scholar 

    15.
    Stocker, T. F. et al (eds). Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change 153 (2013).

    16.
    Otto, O. et al. Energy budget constraints on climate response. Nat. Geosc. 6, 415–416 (2013).
    ADS  CAS  Article  Google Scholar 

    17.
    Fyfe, J. C., Gillett, N. P. & Zwiers, F. W. Overestimated global warming over the past 20 years. Nat. Clim. Change 3, 767–769 (2013).
    ADS  Article  Google Scholar 

    18.
    Kosaka, Y. & Xie, S.-P. Recent global-warming hiatus tied to equatorial Pacific surface cooling. Nature 501, 403–407 (2013).
    ADS  CAS  PubMed  Article  Google Scholar 

    19.
    Santer et al. Volcanic contribution to decadal changes in tropospheric temperature. Nat. Geosc. 7, 185–189 (2014).
    ADS  CAS  Article  Google Scholar 

    20.
    Trenberth, K. & Fasullo, J. An apparent hiatus in global warming? Earth’s Future 1, 19–32 (2013).
    ADS  Article  Google Scholar 

    21.
    Smith, D. Has global warming stalled? Nat. Clim. Change 3, 618–619 (2013).
    ADS  Article  Google Scholar 

    22.
    Guemas, V., Doblas-Reyes, F. J., Andreu-Burillo, I. & Asif, M. Retrospective prediction of the global warming slowdown in the past decade. Nat. Clim. Change 3, 649–653 (2013).
    ADS  Article  Google Scholar 

    23.
    Chen, X. & Tung, K.-K. Varying planetary heat sink led to global-warming slowdown and acceleration. Science 345, 897–903 (2014).
    ADS  CAS  PubMed  Article  Google Scholar 

    24.
    Boykoff, M. Media discourse on the climate slowdown. Nat. Clim. Change 4, 156–158 (2014).
    ADS  Article  Google Scholar 

    25.
    Hawkins, E., Edwards, T. & McNeall, D. Pause for thought. Nat. Clim. Change 4, 154–156 (2014).
    ADS  Article  Google Scholar 

    26.
    England, M. H. et al. Recent intensification of wind-driven circulation in the pacific and the ongoing warming hiatus. Nat. Clim. Change 4, 222–227 (2014).
    ADS  Article  Google Scholar 

    27.
    Karl, T. R. et al. Possible artifacts of data biases in the recent global surface warming hiatus. Science 348, 1469–1472 (2015).
    ADS  CAS  PubMed  Article  Google Scholar 

    28.
    Cowtan et al. Robust comparison of climate models with observations using blended land air and ocean sea surface temperatures. Geophys. Res. Lett. 42, https://doi.org/10.1002/2015GL064888 (2015).

    29.
    Cohen, J. L., Furtado, J. C., Barlow, M., Alexeev, V. A. & Cherry, J. E. Asymmetric seasonal temperature trends. Geophys. Res. Lett. 39, L22705 (2012).
    Google Scholar 

    30.
    Medhaug, I., Stolpe, M. B., Fischer, E. M. & Knutti, R. Reconciling controversies about the ‘global warming hiatus’. Nature 545, 41–47 (2017).
    ADS  CAS  PubMed  Article  Google Scholar 

    31.
    Balmaseda, M. A., Trenberth, K. E. & Källén, E. Distinctive climate signals in reanalysis of global ocean heat content. Geophys. Res. Lett. 40, 1754–175928 (2013).
    ADS  Article  Google Scholar 

    32.
    Wills, R. C., Schneider, T., Wallace, J. M., Battisti, D. S. & Hartmann, D. L. Disentangling global warming, multidecadal variability, and El Niño in Pacific temperatures. Geophys. Res. Lett. 45, 2487–249 (2018).
    ADS  Article  Google Scholar 

    33.
    Cai, W. et al. Pantropical climate interactions. Science 363, eaav4236 (2019).

    34.
    Aregawi, M. et al. Time series analysis of trends in malaria cases and deaths at hospitals and the effect of antimalarial interventions, 2001–2011, Ethiopia. PLoS ONE 9, e106359 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    35.
    Taffese, H. S. et al. Malaria epidemiology and interventions in Ethiopia from 2001 to 2016. Infect. Dis. Poverty 7, 103 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    36.
    Vautard, R., Yiou, P. & Ghil, M. Singular-spectrum analysis: a toolkit for short, noisy chaotic signals. Phys. D 58, 95–126 (1992).
    Article  Google Scholar 

    37.
    Ghil, M. et al. Advanced spectral methods for climatic time series. Rev. Geophys. 40, 1–41 (2002).
    Article  Google Scholar 

    38.
    Harris, T. J. & Yuan, H. Filtering and frequency interpretations of singular spectrum analysis. Phys. D 239, 1958–1967 (2010).
    MathSciNet  CAS  MATH  Article  Google Scholar 

    39.
    Anyamba, A., Tucker, C. J. & Eastman, J. R. NDVI anomaly patterns over Africa during the 1997/1998 ENSO warm event. Int. J. Rem. Sens. 22, 1847–1859 (2001).
    Article  Google Scholar 

    40.
    Nicholson, S. E. & Kim, J. The relationship of the El Niño Southern oscillation to African rainfall. Int. J. Climatol. 17, 117–135 (1997).
    Article  Google Scholar 

    41.
    Reason, C. J. C. & Rouault, M. ENSO-like decadal variability and South African rainfall. Geophys. Res. Lett. 29, 1638 (2002).
    ADS  Article  Google Scholar 

    42.
    Rodó, X. Reversal of three global atmospheric fields linking changes in SST anomalies in the Pacific, Atlantic and Indian ocean at tropical latitudes and midlatitudes. Clim. Dyn. 18, 203–217 (2001).
    Article  Google Scholar 

    43.
    Rodó, X., Pascual, M., Fuchs, G. & Faruque, A. S. G. ENSO and cholera: A nonstationary link related to climate change? Proc. Natl Acad. Sci. USA 99, 12901–12906 (2002).
    ADS  PubMed  Article  CAS  Google Scholar 

    44.
    Saji, N. et al. A dipole mode in the tropical Indian Ocean. Nature 401, 360–363 (1999).
    ADS  CAS  PubMed  Google Scholar 

    45.
    Neale, R. B. et al. Description of the NCAR Community Atmosphere Model (CAM 5.0), NCAR/TN-486+STR. http://www.cesm.ucar.edu/models/cesm1.0/cam/docs/description/cam5_desc.pdf (2012).

    46.
    Shanks, D., Hay, S., Stern, D., Biomndo, K. & Snow, R. Meteorologic Influences on Plasmodium falciparum malaria in the highland tea estates of Kericho, Western Kenya. Emerg. Infect. Dis. 12, 1404–1408 (2002).
    Article  Google Scholar 

    47.
    Negash, et al. Malaria Epidemics in the Highlands of Ethiopia. East Afr. Med. J. 82, https://doi.org/10.4314/eamj.v82i4.9279 (2005).

    48.
    Taffese, H. S. et al. Malaria epidemiology and interventions in Ethiopia from 2001 to 2016. Infect. Dis. Poverty 7, 103 (2018).

    49.
    Fetene, et al. The Ethiopian health extension program and variation in health systems performance: what matters? PLoS ONE 11, e0156438 (2016).

    50.
    PMI, Presidents Malaria Initiative. Ethiopia, Malaria Operational Plan FY, 2018. https://www.pmi.gov/docs/default-source/default-document-library/malaria-operational-plans/fy-2018/fy-2018-ethiopia-malaria-operational-plan.pdf?sfvrsn=5 (2018).

    51.
    Aregawi, et al. Time series analysis of trends in malaria cases and deaths at hospitals and the effect of antimalarial interventions, 2001–2011, Ethiopia. PLoS ONE 9, e106359 (2014).

    52.
    Roy, M., Bouma, M. J., Ionides, E. L., Dhiman, R. C. & Pascual, M. The potential elimination of Plasmodium vivax malaria by relapse treatment: insights from a transmission model and surveillance data from NW India. PLoS Negl. Trop. Dis. 7, 1–10 (2013).
    Article  Google Scholar 

    53.
    Rhein, M., Rintoul, S. R., & Aoki, S. in Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of The Intergovernmental Panel on Climate Change (2013).

    54.
    Roxy, M. K., Ritika, K., Terray, P. & Masson, S. The curious case of Indian Ocean Warming. J. Clim. 27, 8501–8509 (2014).
    ADS  Article  Google Scholar 

    55.
    Diro, G. T. DiroD., Grimes, I. F. & Black, E. Teleconnections between Ethiopian summer rainfall and sea surface temperature: Part I-observation and modelling. Clim. Dyn. 37, 103–119 (2010).
    Article  Google Scholar 

    56.
    Beltrando, G. & Camberlin, P. Interannual variability of rainfall in the eastern horn of Africa and indicators of atmospheric circulation. Int. J. Climatol. 13, 533–546 (1993).
    Article  Google Scholar 

    57.
    Gissila, T., Black, E., Grimes, E., & Slingo, J. M. Seasonal forecasting of the Ethiopian Summer rains. Int. J. Climatol. 24, https://doi.org/10.1002/joc.1078. (2004).

    58.
    Hansen, J., Sato, M., Kharecha, P. & von Schuckmann, K. Earth’s energy imbalance and implications. Atmos. Chem. Phys. 11, 13421–13449 (2011).
    ADS  CAS  Article  Google Scholar 

    59.
    Korecha, D. & Barnston, A. G. Predictability of June-September Rainfall in Ethiopia. Monthly Weather Rev. 135, 628–650 (2007).
    ADS  Article  Google Scholar 

    60.
    Funk, C. et al. Warming of the Indian Ocean threatens eastern and southern African food security but could be mitigated by agricultural development. Proc. Nat. Acad. Sci. USA 105, 11081–11086 (2008).
    ADS  CAS  PubMed  Article  Google Scholar 

    61.
    Williams, A. P. & Funk, C. A. Westward extension of the warm pool leads to a westward extension of the Walker circulation, drying eastern Africa. Clim. Dyn. 37, 2417–2435 (2011).
    Article  Google Scholar 

    62.
    Hoell, A., Hoerling, M., Eischeid, Quan, X., & Liebmann, B. Reconciling Theories for Human and Natural Attribution of Recent East Africa Drying. J. Clim. 30, https://doi.org/10.1175/JCLI-D-16-0558.1. (2016).

    63.
    Kucharski, F., Kang, I. S., Farneti, R. & Feudale, L. Tropical Pacific response to 20th century Atlantic warming. Geophys. Res. Lett. 38, L03702 (2011).
    ADS  Article  Google Scholar 

    64.
    Kug, J.-S. & Kang, I.-S. Interactive feedback between ENSO and the Indian Ocean. J. Clim. 19, 1784–1801 (2006).
    ADS  Article  Google Scholar 

    65.
    Luo, J.-J., Sasaki, W. & Masumoto, Y. Indian Ocean warming modulates Pacific climate change. Proc. Natl Acad. Sci. USA 109, 18 701–18 706 (2012).
    CAS  Article  Google Scholar 

    66.
    Meyrowitsch, D. W. et al. Is the current decline in malaria burden in sub-Saharan Africa due to a decrease in vector population? Malar. J. 10, 188 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    67.
    Baeza, A., Bouma, M. J., Dhiman, R. & Pascual., M. Malaria control under unstable dynamics: reactive vs. climate-based strategies. Acta Trop. Spec. Sect. Hum. Infect. Dis. Environ. Chang. 129, 42–51 (2014).
    Google Scholar 

    68.
    Harris, I., Jones, P. D., Osborn, T. J. & Lister, D. H. Updated high-resolution grids of monthly climatic observations – the CRU TS3.10 Dataset. Int. J. Climatol. 34, 623–642 (2014).
    Article  Google Scholar 

    69.
    Kalnay, E. et al. The NCEP/NCAR 40-year reanalysis project. Bull. Am. Met. Soc. 77, 437–471 (1996).
    ADS  Article  Google Scholar 

    70.
    Rayner, N. A. et al. Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res. 108, 4407 (2003).
    Article  Google Scholar 

    71.
    Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M. & R., C. Francis A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Am. Met. Soc. 78, 1069–1079 (1997).
    ADS  Article  Google Scholar 

    72.
    Thomson, D. J. Spectrum estimation and harmonic analysis. Proc. IEEE 70, 1055–1096 (1982).
    ADS  Article  Google Scholar 

    73.
    Percival, D. B., and Walden, A. T. Spectral Analysis for Physical Applications: Multitaper and Conventional Univariate Techniques (1993).

    74.
    Press, W. H., Teukolsky, S. A., Vetterling, W. T. & Flannery, B. P. Section 13.4.3. Multitaper Methods and Slepian Functions, Numerical Recipes: The Art of Scientific Computing 3rd edn. (2007).

    75.
    Rodríguez-Arias, M. A. & Rodó, X. A primer on the study of transitory dynamics in ecological series using the scale-dependent correlation analysis. Oecologia 138, 485–504 (2004).
    ADS  PubMed  Article  Google Scholar 

    76.
    Rodó, X. & Rodríguez-Arias, M. A. A new method to detect transitory signatures and local time/space variability structures in the climate system: the scale-dependent correlation analysis. Clim. Dyn. 27, 441–458 (2006).
    Article  Google Scholar 

    77.
    Laneri, K. et al. Forcing versus feedback: epidemic malaria and monsoon rains in northwest India. PLoS Comput. Biol. 6, e1000898 (2010).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    78.
    Laneri, K. et al. Dynamical malaria models reveal how immunity buffers effect of climate variability. Proc. Nat. Acad. Sci. USA 112, 8786–8791 (2015).
    ADS  CAS  PubMed  Article  Google Scholar 

    79.
    Roy, M., Bouma, M., Dhiman, R. C. & Pascual, M. Predictability of epidemic malaria under non-stationary conditions with process-based models combining epidemiological updates and climate variability. Malar. J. 14, 1 (2015).
    Article  CAS  Google Scholar 

    80.
    Ionides, E., Bretó, C. & King, A. Inference for nonlinear dynamical systems. Proc. Natl Acad. Sci. USA 103, 18438–18443 (2006).
    ADS  CAS  PubMed  Article  Google Scholar 

    81.
    King, A., Nguyen, D. & Ionides, E. Statistical inference for partially observed Markov processes via the R package pomp. J. Stat. Softw. 69, 1–43 (2016).
    Article  Google Scholar 

    82.
    Hurrell, J. W., Hack, J., Shea, D., Caron, J. & Rosinski, J. A new sea surface temperature and sea ice boundary dataset for the community atmosphere model. J. Clim. 21, 5145–5153 (2008).
    ADS  Article  Google Scholar 

    83.
    Meinshausen, M. et al. The RCP greenhouse gas concentrations and their extension from 1765 to 2300. Clim. Change https://doi.org/10.1007/s10584-011-0156-z (2011).

    84.
    Cionni et al. Ozone database in support of CMIP5 simulations: results and corresponding radiative forcing. Atmos. Chem. Phys. 11, 11267–11292 (2011).
    ADS  CAS  Article  Google Scholar 

    85.
    Lamarque et al. CAM-chem: description and evaluation of interactive atmospheric chemistry in the Community Earth System Model. Geophys. Mod. Dev. 5, 369–411 (2012).
    Article  Google Scholar 

    86.
    Danielson, J. J. & Gesch, D. B. Global multi-resolution terrain elevation data 2010 (GMTED2010). Open-file report 2011-1073 (2011). More

  • in

    Growing support for valuing ecosystems will help conserve the planet

    The Sierra de Manantlán biosphere reserve in Mexico is a source of clean water for urban residents in nearby cities.Credit: Adriana Margarita Larios Arellano/Shutterstock

    Sierra de Manantlán is a 140,000-hectare biosphere reserve in west central Mexico. It is home to 3,000 plant species and a forest whose soils and limestone mountains enable purified water to reach the nearby town of Colima.
    Twenty years ago, researchers at the University of Guadalajara in Mexico proposed that Colima should consider paying to use the forest’s clean water, and that the money could go to supporting the biosphere reserve’s inhabitants.
    The 30,000 people who lived in the forest were poor and in ill health. Unemployment was high, and there were few schools or medical clinics. But the absence of buildings, piped water and electric power had an unintended consequence: it was keeping the forest intact. In return for looking after nature, the researchers argued, the people of Sierra de Manantlán should be compensated, and the funds used for education, health care and employment training. Although not a new idea for Mexico, it was rejected by the city’s authorities. The concept that a forest ecosystem had monetary value — and that its custodians could be compensated — was controversial and much misunderstood.

    Last week, however, countries took a giant step towards enabling public authorities to put a value on their environment. At its annual meeting, the United Nations Statistical Commission — whose members are responsible for setting and verifying standards for official statistics in their countries — laid out a set of principles for measuring ecosystem health and calculating a monetary value. These principles, known as the System of Environmental-Economic Accounting Ecosystem Accounting (SEEA EA), are set to be adopted by many countries on 11 March.
    The principles were agreed after a 3-year writing and review process that involved 100 experts and 500 reviewers from various disciplines and countries. Once adopted, they will give national statisticians an internationally agreed rule book. It will provide a template for payments for ecosystem services — such as those once proposed for Colima — and an official benchmark against which the condition of ecosystems can be judged by policymakers and researchers over time.
    The decision didn’t go as far as it might have done. The overwhelming majority of participating countries — led by Brazil, Colombia, India, Mexico and South Africa, among others — wanted the new rules to be designated as a statistical standard. These countries, rich in biodiversity, want to get on with valuing their natural systems, partly so that any ecological losses can be compared with potential gains from economic development. The designation of a statistical standard would also have enabled statistics offices to access public and international funding to carry out what would be regarded as a core part of their work, and not something voluntary or non-essential.
    But the United States and a number of European Union countries objected. This was partly on the grounds that there is still much debate over valuation methodology, meaning that it is too soon to use ‘standard’ as a label. This setback was unfortunate: participating countries could have adopted the label while creating a system for revision and refinement, ensuring that the new standard could continue to be improved. Fortunately, the meeting’s attendees chose the next best thing — calling the rules “internationally recognized statistical principles and recommendations”.

    The objections raised are a reminder that opinions on setting monetary values for nature are deeply held, with persuasive arguments on all sides. Some argue that nature is too valuable to be regarded in the same way as a commodity, and belongs to all. Valuation in the economic sense suggests that someone has ownership rights — but ecosystem services are rarely, if ever, ‘owned’ by anyone. The new principles do take this into account.
    The record of the statisticians’ meeting shows that much debate remains on how to value something that isn’t bought and sold in a conventional way. But at the same time, this is an active area of research. Many studies have been captured in a landmark report, The Economics of Biodiversity: The Dasgupta Review, published last month by the UK Treasury. The Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services is also conducting a review of the concept of valuation, which will include additional perspectives from the humanities, and voices from under-represented communities, such as Indigenous peoples.
    The debates will continue, but agreement between the world’s statisticians is nevertheless an important step. It means, for example, that those wishing to compensate low-income and marginalized communities for protecting nature — such as the communities in Sierra de Manantlán — now have an internationally agreed template to work from. And policymakers will have to contend with the heads of statistics agencies if they object. UN chief economist Elliot Harris rightly called the new principles a game changer. “The economy needs a bailout, but so does nature,” he said. “What we measure, we value, and what we value, we manage.” Momentum on valuing ecosystem services is now unstoppable, and that is a good thing. More

  • in

    Bioinformatic analysis of chromatin organization and biased expression of duplicated genes between two poplars with a common whole-genome duplication

    An improved reference genome of P. alba var. pyramidalis
    To identify the major structural variation between the genomes of these two species, we first produced a chromosome-level genome assembly of P. alba var. pyramidalis using single-molecule sequencing and chromosome conformation capture (Hi-C) technologies, and then performed comparative genomic analysis with a recently published genome assembly of P. euphratica37. The resulting assembly of P. alba var. pyramidalis consisted of 131 contigs spanning 408.08 Mb, 94.74% (386.61 Mb) of which were anchored onto 19 chromosomes (Supplementary Fig. S1 and Supplementary Tables S1–S3). A total of 40,215 protein-coding genes were identified in this assembly (Supplementary Table S4). The content of repetitive elements in the genome of P. alba var. pyramidalis (138.17 Mb, 33.86% of the genome) is 188.94 Mb less than that of P. euphratica (327.11 Mb, 56.95% of the genome), which contributes greatly to their differences in genome size (Supplementary Table S5).
    3D organization of the poplar genomes
    To characterize the spatial organization and evolution of poplar 3D genomes at a high resolution, we performed Hi-C experiments using HindIII for P. euphratica and P. alba var. pyramidalis, generating a total of 482.95 million sequencing read pairs. These data were mapped to their respective reference genome sequences. After stringent filtering, 81.72 and 94.61 million usable valid read pairs were obtained in P. euphratica and P. alba var. pyramidalis, respectively, and used for subsequent comparative 3D genome analysis (Supplementary Table S6). In addition, we profiled the DNA methylation and transcriptomes of the same tissue samples to provide a framework for understanding the relationships among epigenetic features and 3D chromatin architecture in poplar.
    We first examined genome packing at the chromosomal level with a genome-wide Hi-C map at 50 kb binning resolution for P. euphratica and P. alba var. pyramidalis. As expected, the normalized Hi-C map from both species showed intense signals on the main diagonal (Fig. 1, and Supplementary Figs. S2 and S3) and a rapid decrease in the frequency of intrachromosomal interactions with increasing genomic distance, indicating frequent interactions between sequences close to each other in the linear genome (Supplementary Fig. S4). Strong intrachromosomal and interchromosomal interactions were also observed on the chromosome arms, implying the presence of chromosome territories in the nucleus, in which each chromosome occupies a limited, exclusive nuclear space16,38.
    Fig. 1: Hi-C heatmaps with compartment region analysis results at 50-kb resolution of P. euphratica chromosome 1 (left) and P. alba var. pyramidalis chromosome 1 (right).

    The heatmaps at the top are Hi-C contact maps at 50-kb resolution, which show global patterns of chromatin interaction in the chromosome. The chromosome is shown from top to bottom and left to right. The ICE-normalized interaction intensity is shown on the color scale on the right side of the heatmap. The track below the Hi-C heatmap shows the partition of A (red histogram, PC1  > 0) and B (green histogram, PC1 5 kb) structural variants ranging from 5 to 446 kb in length in the alignment of the two genomes, including 719 inversions, 476 translocations, and 7947 and 10,093 unique regions in P. alba var. pyramidalis and P. euphratica, respectively (Supplementary Tables S10 and S11).
    To characterize the relationship between structural variation and spatial organization of the poplar genomes, we first analyzed the conservation of A/B compartments between P. alba var. pyramidalis and P. euphratica, using a 50-kb Hi-C matrix. The results showed that 71.52% (145.75 Mb in P. euphratica and 145.63 Mb in P. alba var. pyramidalis) of the total length of the syntenic regions have the same compartment status between the two species, while 43.68 and 43.71 Mb of the genomic regions exhibit A/B compartment switching in P. alba var. pyramidalis and P. euphratica, respectively (Fig. 3a). For the regions with structural variation, we found that 77% of the inversion events between the two genomes had no effects on their compartment status, while 61% of the translocation events occurred within the regions exhibiting compartment switching (Fig. 4a and Supplementary Table S10). Moreover, we also found that 38.59% and 33.39% of the nonsyntenic regions were identified as A compartments in P. alba var. pyramidalis and P. euphratica, respectively, indicating that the large-scale insertions and/or deletions are biased to occur at heterochromatic regions (Fig. 4b). We further assessed the conservation of genome organization at the TAD level by examining whether the orthologous genes within the same TAD in one species could still be located within the TAD in another species19,21,23. The results indicated that only 48.04% of TADs from P. alba var. pyramidalis and 40.95% from P. euphratica were substantially shared between the two species (Figs. 3b, c). Taken together, these results indicated that the 3D genome organization shows surprisingly low conservation across poplar species at both the compartmental and TAD levels.
    Fig. 3: Evolutionary conservation of compartment status and TADs across P. euphratica and P. alba var. pyramidalis.

    a Overlap of compartment status between syntenic regions in P. euphratica and P. alba var. pyramidalis. b Overlap of TADs between syntenic regions in P. euphratica and P. alba var. pyramidalis. c Example of conserved TAD structures across a syntenic region between P. euphratica and P. alba var. pyramidalis. The TADs are outlined by black triangles in the heatmaps, and the position of the TAD domains is indicated by alternating blue-green line segments. The mean cf value used to identify the domains is also shown. The orthology tracks of these conserved domains are shown at the bottom

    Full size image

    Fig. 4: Relationship between structural variation and spatial organization of the genomes of P. euphratica and P. alba var. pyramidalis.

    a Analysis of compartment inversion (left) and translocation (right) across P. euphratica and P. alba var. pyramidalis. b Analysis of compartments of species-specific regions in P. euphratica (left) and P. alba var. pyramidalis (right)

    Full size image

    Relationship between chromatin interactions and expression divergence of WGD-derived paralogs
    Poplar species have undergone a recent WGD event followed by diploidization, a process of genome fractionation that leads to functional and expression divergence of the duplicated gene pairs27,28,33. Although no biased gene loss or expression dominance was found between the two poplar subgenomes, there is evidence that nearly half of the WGD-derived paralogs have diverged in expression32,33. To explore the potential role of chromatin dynamics on the observed expression patterns of duplicated genes, we examined their differences in chromatin interaction patterns for both species. We first identified a total of 10,438 and 9754 paralogous gene pairs showing interchromosomal interactions in P. euphratica and P. alba var. pyramidalis, respectively. After correlating the frequency of chromatin interactions with their differences in expression, we found that gene pairs with biased expression (more than twofold differences in expression levels) interacted less frequently than gene pairs with similar expression levels in both species (P = 1.71 × 10−6 and 7.20 × 10−7 for P. euphratica and P. alba var. pyramidalis, respectively, Mann–Whitney U test; Fig. 5a). We also estimated the interaction score (the average of the distance-normalized interaction frequencies) for bins involved in the paralogous gene pairs and quantified their differences in interaction strength (Supplementary Fig. S7 and Supplementary Table S12)3,23. Our results showed that for gene pairs with biased expression, highly expressed gene copies have stronger interaction strengths than weakly expressed copies (P = 2.10 × 10−12 and 2.74 × 10−2 for P. alba var. pyramidalis and P. euphratica, respectively, Mann–Whitney U test), while no significant differences were observed for gene pairs with similar expression levels (Fig. 5b). We further investigated these phenomena at the level of high-order chromatin architecture and found that the gene pairs located in conserved TADs had similar expression levels (P = 2.68 × 10−3 and 7.86 × 10−6 for P. euphratica and P. alba var. pyramidalis, respectively, Mann–Whitney U test; Supplementary Fig. S8). Overall, our analyses indicate that the extensive expression divergence between WGD-derived paralogs in Populus is associated with the differences in their chromatin dynamics and 3D genome organization, and suggest that this organization may function as a key regulatory layer underlying expression divergence during diploidization.
    Fig. 5: Comparison of interaction levels between WGD-derived paralogs with biased/similar expression in P. euphratica and P. alba var. pyramidalis.

    a The box plot shows that the interaction frequency of WGD-derived paralogs with biased (fold change  > 2) and similar (fold change  More